You are on page 1of 15

3

Rheological Characterization
Annika Björn, Paula Segura de La Monja,
Anna Karlsson, Jörgen Ejlertsson and Bo H. Svensson
Department of Thematic Studies, Water and Environmental Studies,
Linköping University,
Sweden

1. Introduction
The biogas process has long been a part of our biotechnical solutions for the handling of
sewage sludge and waste. However, in many cases the existing process applications need to
be optimized to improve the extent of biogas production as a part of the energy supply in a
sustainable and viable society. Although the principles are well known, process
disturbances and poor substrate utilization in existing biogas plants are common and are in
many cases likely linked to changes in the substrate composition.
Changes in substrate composition can be done as a means to obtain a more efficient
utilization of existing biogas facilities, which today treating mainly manure or sewage
sludge. By bring in more energy rich residues and wastes a co-digestion process with higher
biogas potential per m3 volatile solids (VS) can often be obtained. However, new and
changing feedstocks may result in shift in viscosity of the process liquid and, hence,
problems with inadequate mixing, break down of stirrers and foaming. These disturbances
may seriously affect the degradation efficiency and, hence, also the gas-production per unit
organic material digested. In turn, operational malfunctions will cause significant logistic
problems and increased operational costs. Changes of the substrate profile for a biogas plant
may also infer modifications of the downstream treatment of the digestate.
Together with high digestion efficiency, i.e. maximum methane formation per reactor
volume and time, the economy of a biogas plant operation depends on the energy invested
to run the process. A main part of the energy consumed during operation of continuous
stirred tank reactors (CSTRs) is due to the mixing of the reactor material (Nordberg and
Edström, 2005). The shear force needed is dependent on the viscosity of the reactor liquid,
where increasing viscosity demands a higher energy input. Active stirring must be
implemented in order to bring the microorganisms in contact with the new feedstock, to
facilitate the upflow of gas bubbles and to maintain an even temperature distribution in the
digester. Up to 90% of biogas CSTR plants use mechanical stirring equipment (Weiland et
al., 2010).
In this context the rheological status of the reactor liquid as well as of the residual digestate
are important for process mixing design and dimensioning. In addition experiences on
rhelogical characterisation of sewage sludge revealing their dependence on the suspended

www.intechopen.com
64 Biogas

solid concentration and on the characteristics of the organic material as well as on the
interactions between particles and molecules in the solution (Foster, 2002). Therefore, this
type of characterisation can be important in process monitoring and control.
The aim of this chapter is to briefly introduce the area of rheology and to present important
parameters for rheological characterization of biogas reactor fluids. Examples are given
from investigations on such parameters for lab-scale reactors digesting different substrates.

2. Rheology
Rheology describes the deformation of a body under the influence of stress. The nature of
the deformation depends on the body’s material conditions (Goodwin & Hughes, 2000).
Ideal solids deform elastically, which means that the solid will deform and then return to its
previous state once the force ceases. In this case, the energy needed for deformation will
mainly be recovered after the stress terminates. If the same force is applied to ideal fluids, it
will make them flow and the energy utilized will disperse within the fluid as heat. Thus, the
energy will not be recovered once the forcing stress is terminated (Goodwin & Hughes,
2000).
For fluids a flow curve or rheogram is used to describe rheological properties. These
properties may be of importance in anaerobic digestion for the dimensioning of e.g. feeding,
pumping and stirring. Rheograms are constructed by plotting shear stress (τ) as a function
of the shear rate () (Tixier et al., 2003; Guibad et al., 2005).
The stress applied to a body is defined as the force (F) divided by the area (A) over which
this force is acting (Eq. 1). When forces are applied in opposite directions and parallel to the
side of the body it is called shear stress (Goodwin & Hughes, 2000). Shear stress (τ; Pa) is
one of the main parameters studied in rheology, since it is the force per unit area that a fluid
requires to start flowing (Schramm, 2000). The shear rate (; s-1) describes the velocity
gradient (Eq. 2). Hence, shear rate is the speed of a fluid inside the parallel plates generated
when shear stress is applied (Pevere & Guibad, 2005).

τ = F/A = N/m2 = Pa (1)

 = dvx/dy = (m/s)/m (2)

2.1 Newtonian fluids


Ideal fluids (e.g. water, methanol, olive oil and glycerol) perform linearly in rheograms, as

equation (Eq. 3) illustrates the flow behaviour of an ideal liquid (Schramm, 2000), where  is
illustrated for glycerol in figure 1, and are identified as Newtonian fluids. The Newtonian

the viscosity (Pa*s). Dynamic viscosity, also called apparent viscosity, describes a fluid’s
resistance of deformation (Pevere & Guibad, 2005). In terms of rheology it is the relation of

maintains a constant value meaning a linear relationship between  and .


shear stress over the shear rate (Eq. 4). For Newtonian fluids the dynamic viscosity

τ=* (3)

=/ (4)

www.intechopen.com
Rheological Characterization 65

When measuring the dynamic viscosity, the fluid is subjected to a force impact caused by
moving a body in the fluid. Resistance to this movement provides a measure of fluid
viscosity. The dynamic viscosity can be measured using a rotation rheometer. The device
consists of an external fixed cylinder with known radius and an internal cylinder or spindle
with known radius and height. The space between the two cylinders is filled with the fluid
subjected to dynamic viscosity analysis.

130

Pa

110

100

90

80

70

 60

50

40

30

20

10

0
0 100 200 300 400 500 600 700 1/s 800

Shear Rate 
.

Fig. 1. Rheogram – flow curve of glycerol (▲) at 20 °C with a linear relationship between
shear stress (; Pa) and shear rate (; s-1), representing a Newtonian liquid.

2.2 Limit viscosity


Limit viscosity (lim) corresponds to the viscosity of a fluid at the maximum dispersion of
the aggregates under the effect of the shear rate (Tixier & Guibad, 2003). The limit viscosity
is estimated through the rheogram, when the dynamic viscosity becomes linear and
constant. This parameter has been shown to be of great value when studying the rheological
characteristics of sludge, since it determines the level of influence of important factors such
as the total solids fraction (TS; Lotito et al., 1997). TS (%) and volatile solids (VS, % of TS) are
parameters measured in the biogas process in order to control the amount of solids that may
be transformed to methane. Also, Pevere and Guibad (2005) reported that the limit viscosity
was sensitive to the physicochemical characteristics of granular sludge, i.e. it was influenced
by changes in the particle size or the zeta potential.

www.intechopen.com
66 Biogas

2.3 Dynamic yield stress


Yield stress (0) is defined as the force a fluid must be exposed to in order to start flowing. It
reflects the resistance of the fluid structure to deformation or breakdown. Rheograms from
rotational viscometer measurements are used as a means to calculate yield stress. It can also
be obtained by applying rheological mathematical models (section 2.6; Spinosa & Loito
2003). Yield stress is important to consider when mixing reactor materials, since the yield
stress is affecting the physico-chemical characteristics of the fluid and impede flow even at
relative low stresses. This might lead to problems like bulking or uneven distribution of
material in a reactor (Foster, 2002).

2.4 Static yield stress


The static yield stress (s) is the yield stress measured in an undisturbed fluid while dynamic
yield stress is the shear stress a fluid must be exposed to in order to become liquid and start
flowing. The fact that both dynamic yield stress and static yield stress sometimes may
appear is explained by the existence of two different structures of a fluid. One structure is
not receptive to the shear stress and tolerates the dynamic yield stress, while a second
structure (a weak gel structure) is built up after the fluid has been resting a certain period of
time (Yang et al., 2009). When these two structures merge, a greater resistance to flow is
generated translated to the static yield stress.
The formation of the weak gel structure may be a result from chemical interactions among
polysaccharides or between proteins and polysaccharides (Yang et al., 2009). The weak gel
structure is quite vulnerable and, thus easily interrupted by increasing shear rates.

2.5 Non-Newtonian fluids


Non-Newtonian fluids do not show a linear relationship between shear stress and shear
rate. This is due to the complex structure and deformation effects exhibited by the materials
involved in such fluids. The non-Newtonian fluids are however diverse and can be
characterised as e.g. pseudoplastic, viscoplastic, dilatant and thixotropic fluids (Schramm,
2000).

2.5.1 Pseudoplastic fluids


Pseudoplastic fluids become thinner when the shear rate increases, until the viscosity
reaches a plateau of limit viscosity. This behaviour is caused by increasing the shear rate and
the elements suspended in the fluid will follow the direction of the current. There will be a
deformation of fluid structures involving a breaking of aggregates at a certain shear rate and
this will cause a limit in viscosity. For pseudoplastic fluids the viscosity is not affected by
the amount of time the shear stress is applied as these fluids are non-memory materials i.e.
once the force is applied and the structure is affected, the material will not recover its
previous structure (Schramm, 2000). Some examples are corn syrup and ketchup.

2.5.2 Viscoplastic fluids


Viscoplastic fluids, such as e.g. hydrocarbon greases, several asphalts and bitumen, behave
as pseudoplastic fluids upon yield stress. They need a predetermined shear stress in order to

www.intechopen.com
Rheological Characterization 67

start flowing. One type of these, the Bingham plastic, requires the shear stress to exceed a
minimum yield stress value in order to go from high viscosity to low viscosity. After this
change a linear relationship between the shear stress and the shear rate will prevail (Ryan,
2003). Examples of Bingham plastic liquids are blood and some sewage sludge’s.

2.5.3 Dilatant fluids


Dilatant fluids become thicker when agitated, i.e. the viscosity increases proportionally with
the increase of the shear rate. Like for the pseudoplastic fluids the stress duration has no
influence, i.e. when the material is disturbed or the structure destroyed it will not go back to
its previous state. Some examples of shear thickening behaviour are honey, cement and
ceramic suspensions.

2.5.4 Thixotropic fluids


Thixotropic fluids are generally dispersions, which when they are at rest construct an
intermolecular system of forces and turn the fluid into a solid, thus, increasing the viscosity.
In order to overcome these forces and make the fluid turn into a liquid and which may flow,
an external energy strong enough to break the binding forces is needed. Thus, as above a
yield stress is needed. Once the structures are broken, the viscosity is reduced when stirred
until it receives its lowest possible value for a constant shear rate (Schramm, 2000). In
opposite to pseudoplastic and dilatant fluids, the viscosity of thixotrpic fluids is time
dependent: once the stirring has ended and the fluid is at rest, the structure will be rebuilt.
This will inform about the fluid possibilities of being reconstructed. Wastewater and sewage
sludge can be examples of fluids with thixotropic behaviour (Seyssieq & Ferasse, 2003) as
well as paints and soap.

2.6 Rheological mathematical models


There are several rheological mathematical models applied on rheograms in order to
transform them to information on fluid rheological behaviour. For non-Newtonian fluids the
three models presented below are mostly applied (Seyssiecq & Ferasse, 2003).

2.6.1 Herschel Bulkley model


The Herschel Bulkley model is applied on fluids with a non linear behaviour and yield
stress. It is considered as a precise model since its equation has three adjustable parameters,

equation 5, where 0 represents the yield stress.


providing data (Pevere & Guibaud, 2006). The Herschel Bulkley model is expressed in

τ =  0 +  * n (5)

be able to compare -values for different fluids they should have similar flow behaviour
The consistency index parameter () gives an idea of the viscosity of the fluid. However, to

index (n). When the flow behaviour index is close to 1 the fluid´s behaviour tends to pass
from a shear thinning to a shear thickening fluid. When n is above 1, the fluid acts as a shear
thickening fluid. According to Seyssiecq and Ferasse (2003) equation 5 gives fluid behaviour
information as follows:

www.intechopen.com
68 Biogas

0 = 0 & n = 1  Newtonian behaviour


0 > 0 & n = 1  Bingham plastic behaviour
0 = 0 & n < 1  Pseudoplastic behaviour
0 = 0 & n > 1  Dilatant behaviour

2.6.2 Ostwald model


The Ostwald model (Eq. 6), also known as the Power Law model, is applied to shear
thinning fluids which do not present a yield stress (Pevere et al., 2006). The n-value in
equation 6 gives fluid behaviour information according to:

τ =  * (n-1) (6)
n < 1  Pseudoplastic behaviour
n = 1  Newtonian behaviour
n > 1  Dilatant behaviour

2.6.3 Bingham model


The Bingham model (Eq. 7) describes the flow curve of a material with a yield stress and a
constant viscosity at stresses above the yield stress (i.e. a pseudo-Newtonian fluid
behaviour; Seyssiecq & Ferasse, 2003). The yield stress (0) is the shear stress () at shear rate
() zero and the viscosity () is the slope of the curve at stresses above the yield stress.

 = 0 +  *  (7)
0 = 0  Newtonian behaviour
0 > 1  Bingham plastic behaviour

3. Rheological characterization of biogas reactor fluids


When considering the rheology for biogas reactors their viscosity is estimated to correspond
to a given TS of the reactor fluid. This is mainly based on historically rheological data from
sewage sludge with known TS values. However, problems may arise when using these TS
relationships for other types of substrates which may impose other rheological
characteristics of the reactor fluids. Furthermore, often low consideration is given to possible
viscosity changes due to variation in feedstock composition etc.
Shift in the viscosity and elasticity properties of the reactor material related to substrate
composition changes can alter the prerequisites for the process regarding mixing (dimension
of stirrers, pumps etc. or reactor liquid circulation) and likely also foaming problems
(Nordberg & Edström, 2005; Menéndez et al., 2006). It may also call for changes in the post
treatment requirements and end use quality of the organic residue e.g. dewatering ability,
pumping and spreading on arable land (Baudez & Coussot, 2001). The additions of enzymes
can be used to reduce the viscosity of the substrate mixture in the digester significantly and
avoid the formation of floating layers (Weiland, 2010; Morgavi et al., 2001). All these factors
affect the total economy for a biogas plant.
In this context differences in the rheological characteristics of biogas reactor fluids as
depending on substrate composition were analyzed and used as examples in this presentation.

www.intechopen.com
Rheological Characterization 69

3.1 Rheological measurments


A rotational rheometer RheolabQC coupled with Rheoplus software (Anton Paar) was used
for different reactor fluids, which recorded the rheograms´ and allowed subsequent data
analysis. The temperature was maintained constant at 370.2 °C. The reactor fluid volume
used for each measurement was 17 ml. Reactor fluids from mesophilic (37°C) lab-scale
reactors (4 L running volume), with a hydraulic retention time (HRT) of 20 days, were
sampled.
Five lab-scale reactors (A-E) were sampled before the daily feeding of substrates. All
reactors had been running for at least three HRTs prior to sampling. The different substrates
treated were slaughter household waste, biosludge from pulp- and paper mill industries,
wheat stillage and cereal residues. The TS values ranged between 3.1−3.9 % for four of the
reactors while one was at 7.7 % (Table 1).

Reactor Digested substrate TS (%)


A Slaughter house waste 3.9
B Biosludge from pulp- and paper mill industry 1 3.8
C Biosludge from pulp- and paper mill industry 2 3.7
D Wheat stillage 3.0
E Cereal residues 7.7
Table 1. Fluids from five lab-scale reactors were chosen for rheological measurements. A
short description of their TS values and substrates are presented.

Rheological measurements were carried out with a three-step protocol where (1) the shear
rate increased linearly from 0 to 800 s-1 in 800 sec., (2) maintaining constant shear rate at 800
s-1 in 30 sec, (3) decreasing linearly the shear rate from 800 to 0 s-1 in 800 sec., according to
Björn et al. (2010). For each sample three measurements were carried out and performed
immediately after sampling or stored at +4 ºC pending analysis.
The fluid behaviour was interpreted by the flow- and viscosity curves according to
Schramm (2000), and the dynamic viscosity, limit viscosity and yield stress were noticed.
The three most common mathematical models for non-Newtonian fluids; Herschel Bulkley
model; Ostwald model (Power Law) and Bingham model, were applied in order to
transform rheogram data values to the rheological behaviour of the fluids. Flow behaviour
index (n) and consistency index (K) were studied.

3.2 Flow and viscosity behaviour characteristics


The flow curves for reactor fluids A-E (Figures 2−3) indicated different flow behaviour
according to the definitions by Schramm (2000). A Newtonian behaviour of reactors A
and D, fed with slaughter house waste and wheat stillage, respectively, was illustrated
where the exerted shear stress was almost proportional to the induced shear rate.
However, a small yield stress of 0.2 Pa and 0.3 Pa were detected, indicating a pseudo-
Newtonian behaviour.
Fluids from reactor B, receiving biosludge from paper mill industry 1 as substrate, indicated
an unusual performance at the beginning of the rheogram with decreasing shear stress,
thereafter a linear increase in shear stress. A yield stress of 14 Pa was detected. A space

www.intechopen.com
70 Biogas

between the curves was noticeable when the shear rate increased and afterwards decreased
for reactor B (Fig. 2). This area describes the degree of thixotropy of this fluid, which means
that the increase of this area is related to the amount of energy required to breaking down
the thixotropic structure. Thus, the flow curves obtained with the three-step protocol
indicated a thixotropic behaviour of reactor fluid B.

30
Pa
26
24

22
20

18
16

 14
12

10
8

6
4

2
0
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700 1/s 800

Shear Rate 
.

Fig. 2. Rheogram - flow curves illustrating shear stress (; Pa) vs shear rate (; s-1) for fluids
from reactor A (▲), B (▲) and C (▲) with a three-step protocol.

Reactors C and E revealed viscoplastic behaviours, i.e. a pseudoplastic behaviour with yield
stress. Reactor C, fed with biosludge from paper mill industry 2, showed a yield stress of 4
Pa (Fig. 2), and reactor E, receiving cereal residues, a yield point of 4.5 Pa (Fig. 3). The yield
stress is defined as the force that a fluid must overcome in order to start flowing (Spinosa &
Lotito, 2003). Also for reactor E, a small space between the curves was noticeable when the
shear rate increased and afterwards decreased (Fig. 3). This area difference might indicate
some degree of thixotropy.

Reactor Flow curve behaviour Viscosity curve behaviour


A Newtonian; pseudo-Newtonian Viscoplastic (pseudo-Newtonian)
B Thixotropic Thixotropic
C Viscoplastic Viscoplastic
D Newtonian; pseudo-Newtonian Pseudoplastic or viscoplastic
E Viscoplastic Viscoplastic
Table 2. The flow and viscosity curves for reactor fluids A-E indicated different fluid
behaviour according to Schramm (2000).

www.intechopen.com
Rheological Characterization 71

5
30 10
Pa Pa·s
4
26 10
24
3
22 10

20
2
18 10

16
 
1
10
14
12 0
10
10
8 -1
10
6
4 -2
10
2
-3
0 10
0 100 200 300 400 500 600 700 1/s 800

Shear Rate 
.

Fig. 3. Rheogram - flow and viscosity curves for reactors D (▲; ●) and E (▲; ●) with a three-
step protocol. Flow curves illustrating shear stress (; Pa) vs shear rate (; s-1) and viscosity
curves illustrating dynamic viscosity (; Pa*s) vs shear stress (; s-1).
30 1 000 000
Pa Pa·s
100 000
26
24
10 000
22
20 1 000

18
100
 
16

14
10
12

10 1
8
0,1
6
4
0,01
2
0 0,001
0 50 100 150 200 250 300 350 400 450 500 550 600 650 700 1/s 800

Shear Rate 
.

Fig. 4. Rheogram - flow and viscosity curves for reactors A (▲;●), B (▲;●) and C (▲;●) with
a three-step protocol. Flow curves illustrating shear stress (; Pa) vs shear rate (; s-1) and
viscosity curves illustrating dynamic viscosity (; Pa*s) vs shear stress (; s-1).

www.intechopen.com
72 Biogas

The viscosity curves (Figures 3−4) did almost correspond to the flow curve behaviour for
the investigated biogas reactor fluids (Table 2). Using the scheme by Schramm (2000) the
viscosity curves for reactor A indicated a viscoplastic liquid and for reactor D a viscoplastic
or pseudoplastic liquid. The viscosity initially dropped very quickly for reactor A,
specifically indicating Bingham viscoplastic fluids with pseudo-Newtonian behaviour.
Generally for reactors A−E, the viscosity decreased with increasing shear rate, until it
reached its limit viscosities (Table 3).

TS Dynamic viscosity Limit viscosity


Reactor Treated substrate
(%) (mPa*s) (mPa*s)
A Slaughter house waste 3.9 18 6
B Biosludge paper mill industry 1 3.8 436 8
C Biosludge paper mill industry 2 3.7 267 29
D Wheat stillage 3.0 33 6
E Cereal residues 7.7 443 36

Table 3. The initial dynamic viscosity and the limit viscosity obtained during interval 1 in
the 3-step protocol analysis for each reactor fluid.

The limit viscosities ranged 6−36 mPa*s with the highest value for reactor E (Table 3).
However, the limit viscosity was similar for reactor C and E despite a difference in TS (%).
The dynamic viscosity ranged 18−443 mPa*s for the reactor fluids (Table 3). The reactors A
and D showed lower dynamic viscosity values compared to reactor B, C and E, possibly due
to their pseudo-Newtonian behaviour. Also, there was a difference in dynamic viscosity
between reactor fluids B and C both receiving biosludge with similar TS (%) but coming
from two different paper mill industries in Sweden. Thus, the results demonstrated that
similar TS values do not necessarily correspond to similar dynamic or limit viscosity values.
This contradicts the results presented by Tixier and Guibad (2003), who reported that an
increase in TS for activated sludge corresponded to a higher limit viscosity and higher yield
stress. Nor, did biosludge from two different Swedish pulp- and paper mill industries with
similar TS give similar viscosity values.
Samples from reactor B showed different rheological behaviour depending on when they
were measured. The yield stress and viscosity increased when the reactor fluids had been
stored and resting compared to when analyzed immediately after sampling, indicating
thixotropic behaviour. Figure 5 illustrates how the B reactor fluid after been resting for 48
hours showed a resistance to flow, known as static yield stress. This value (24 Pa) decreased
until it reached the dynamic yield stress (7 Pa), which was the value needed in order to
become liquid and start flowing. When the analysis was done right after the first
measurement (Second measurement), the fluid had already been stirred, so the two different
structures that form the resistance to flow were now mixed and no static yield stress was
detected. The static yield stress might be initiated by several factors, e.g. weakness of the
fluid structure, low mixed liquid solid suspension (MLSS) concentration, small size of
particles and poor dewater ability (Pevere et al. 2006).

www.intechopen.com
Rheological Characterization 73

Fig. 5. Rheogram – shear stress (; Pa) vs shear rate (; s-1) of reactor B. Four measurements
of the same sample were made after different sample resting times (1st▲; 2nd▲; 3rd▲;
4th ▲).

3.3 Mathematical modeling


Also, the Herchel-Bulkley model indicated that reactor fluid A performed as a pseudo-
Newtonian fluid called Bingham plastic, since the yield stress-value was > 0 (0.24 Pa) and a
flow behaviour index of 1.06 (Table 4). Results obtained by the Ostwald and Bingham
models confirmed a Bingham plastic behaviour of reactor A. However, since the 0-value
was almost 0 and the n-value 1 it was also closely performing as a Newtonian fluid which is
consistent with the flow curve appearance (Fig. 2). However, when studying the viscosity
curve (Fig. 4) the results showed an initial viscosity decrease and then a constant viscosity
indicating a pseudo-Newtonian fluid behaviour.
The Herschel-Bulkley and Ostwald models both indicated a pseudoplastic behaviour of
reactor D, since the 0-value was 0 and n < 1 (Table 4). The Bingham model gave a yield
stress of 0.33 Pa which did not indicate Newtonian or Bingham plastic behaviour. Thus, the
common results for reactor D strongest indicate a pseudoplastic fluid behaviour.
Reactor B was hard to define also when modelling the rheogram data values of figure 4. The
regression values were low for all three mathematical models (Table 4). However, the
Herschel-Bulkely model had a flow behaviour index n>1 indicating that the fluid acted as a
shear thickening (dilatant) fluid, but the Ostwald and Bingham models indicated
pseudoplastic and Bingham plastic behaviours, respectively. When the static yield stress
appeared in the reactor B rheogram (Figures 2 and 4), the flow behaviour index showed
shear thickening fluid behaviour (n=3.4) and a limit viscosity of 8 mPa*s. This also

www.intechopen.com
74 Biogas

corresponded to a low consistency value (5*10-10). At the static yield stress of 24 Pa (Fig. 5),
the flow behaviour index showed shear thickening fluid behaviour (n=1.41) and a limit
viscosity of 22 mPa*s. This also corresponded to a low consistency value (5*10-4). As soon as
the fluid was measured again, n decreased (0.70) showing a pseduoplastic behaviour and K
increased (0.11) indicating that the consistency of the reactor material was higher. The limit
viscosity was 17 mPa*s. These results showing a time dependency and structure recovery
strengthen the arguments for a thixotropic fluid behaviour of reactor B. Once the stirring has
ended and the fluid was at rest, the fluid structure starts to rebuild. Therefore, the viscosity
become time dependent. This information is important to consider for biogas reactor
performance, e.g. when applying semi-continuous mixing.

Herschel-Bulkley Ostwald Bingham

0 n  R2 n  R2 0 R2

A 0.24 1.06 0.003 0.93 0.69 0.35 0.84 0.21 0.92


B 2.57 3.40 5*10-10 0.45 0.08 2.28 0.002 1.88 0.12
C 2.89 0.59 0.42 0.99 0.44 1.23 0.99 6.36 0.95
D 0 0.65 0.04 0.88 0.64 0.04 0.87 0.33 0.95
E 2.38 0.49 0.98 0.96 0.39 1.98 0.96 8.31 0.91

reactors A−E. 0: yield stress (Pa); n: flow behaviour index; : Consistency index; R2:
Table 4. The results obtained from mathematical modelling of rheogram data of fluids from

regression coefficient.

Also, fluids from reactor C and E were hard to define from modelling of the rheogram data
because they gave indications for fluids being between pseudoplastic and Bingham plastic
behaviours, i.e. the 0-values were >0 (2.89 and 2.38) and n <1 (Table 4).

4. Conclusion
The biogas reactor fluids investigated were behaving viscoplasticly, since they had yield
stress and one of them was also thixotropic, due to its partial structure recovering. However,
the reactor treating slaughterhouse waste was very close to act as a Newtonian fluid. Also,
there was a difference in dynamic- and limit viscosities depending on the substrates used.
The results demonstrated that similar TS values did not necessarily correspond to similar
flow and viscosity behaviours. Nor, did biosludge from two different Swedish paper mill
industries with similar TS show similar viscosity values.
To encounter problems related to involvement of new substrates and/or co-digestions in
existing facilities, investigations for possible viscosity changes are needed. Ongoing research
will hopefully provide an important basis for predictions of changes in rheology linked to
the composition of the organic materials, which are translated in the process. This is
important in order to achieve proper designs in relation to possible variation in substrate
mixes in conjunction with new constructions, but also to better control material flows in the
existing facilities to avoid disturbances in the reactor performance.

www.intechopen.com
Rheological Characterization 75

5. References
Baudez, J-C. & Coussot, P. (2001). Rheology og aging concentrated, polymeric
suspensions: application to pasty sewage sludges. Jour Rheol, Vol.45, No.5, pp.
1123-1139.
Björn, A. ; Segura de La Monja, P. ; Karlsson, A. ; Ejlertsson, J. & Svensson, BH. (2010).
Differences and shifts in the rheological characteristics of fluids in controlled
stirred tank reactors for biogas production. Poster presentation at World
Conference in Anaerobic Digestion (AD12), Guadalajara, Mexico October 31rd o
November 4th 2010.
Foster, CF. (2002). The rheological and physico-chemical characteristics of sewage sludges.
Enzyme Microb Technol, Vol.30, pp. 340-345.
Goodwin, JW. & Hughes, RW. (2000). Rheology for Chemists, an introduction. Royal Society of
Chemist. Cambridge, UK.
Guibad, G.; Tixier, N. & Baudu, M. (2005). Hysteresis area, a rheological parameter used as a
tool to assess the ability of filamentous sludges to settle. Process Biochem, Vol.40, pp.
2671-2676.
Lotito, V.; Spinosa, L.; Minini, G. & Antonaci, R. (1997). The rheology of sewage sludge at
different steps of treatment. Water Sci Tech, Vol.36, No.11, pp. 79-85.
Menéndez, M. & Paredes, B. (2006). Rheological behaviour and organoleptic effects of
ovalbumin addition in yoghurt mousse production. Jour Dairy Sci, Vol.89, pp.
951-962.
Morgavi, DP.; Beauchemin, KA. & Nsereko, LM. (2001). Resistance of feed enzymes to
proteolytic inactivation by rumen microorganisms and gastrointestinal proteases.
Jour Anim Sci, Vol.79, pp. 1621-1630.
Nordberg, Å. & Edström, M. (2005). Co-digestion of energy crops and the source-sorted
organic fraction of municipal solid waste. Water Sci Tech, Vol.52, No.1-2, pp.
217-222.
Pevere, A. & Guibad, G. (2005). Viscosity evolution of anaerobic granular sludge. Biochem
Engin Jour, pp. 315-322.
Pevere, Å.; Guibad, G; van Hullenbusch, E.; Lens, P. & Baudu, M. (2006). Viscosity evolution
of anaerobic granular sludgee. Biochem Engin Jour, Vol.27, pp. 315.322
Schramm, G. (2000). A practical Approach to Rheology and Rheometry, (2nd Ed.), Thermo Haake
Rheology, Germany
Seyssiecq, I.; Ferasse, J.H. & Roche, N. (2003). State-of-the-art: rheological characterization of
waste water treatment sludge. Biochem Engin Jour, Vol.16, pp. 41-56.
Spinosa, L., & Lotito, V. (2003). A simple method for evaluating yield stress. Adv Environ
Research, Vol.7, pp. 655-659.
Tixier, N.; Guibad, G. & Baudu, M. (2003). Determination of some rheological
parameters for the characterization of activated sludge. Bioresource Tech, Vol.90,
pp. 215-220.
Weiland, P. (2010). Biogas production: current state and perspectives. Appl Microbiol
Biotechnol, Vol.85, pp. 849-860.

www.intechopen.com
76 Biogas

Yang, F. ; Bick, A. & Shandalov, S. (2009). Yield stress and rheological characteristics of
activated sludge in an airlift membrane bioreactor. Jour Membrane Sci, Vol.334, pp.
83-90.

www.intechopen.com
Biogas
Edited by Dr. Sunil Kumar

ISBN 978-953-51-0204-5
Hard cover, 408 pages
Publisher InTech
Published online 14, March, 2012
Published in print edition March, 2012

This book contains research on the chemistry of each step of biogas generation, along with engineering
principles and practices, feasibility of biogas production in processing technologies, especially anaerobic
digestion of waste and gas production system, its modeling, kinetics along with other associated aspects,
utilization and purification of biogas, economy and energy issues, pipe design for biogas energy,
microbiological aspects, phyto-fermentation, biogas plant constructions, assessment of ecological potential,
biogas generation from sludge, rheological characterization, etc.

How to reference
In order to correctly reference this scholarly work, feel free to copy and paste the following:

Annika Björn, Paula Segura de La Monja, Anna Karlsson, Jörgen Ejlertsson and Bo H. Svensson (2012).
Rheological Characterization, Biogas, Dr. Sunil Kumar (Ed.), ISBN: 978-953-51-0204-5, InTech, Available
from: http://www.intechopen.com/books/biogas/rheological-characterization

InTech Europe InTech China


University Campus STeP Ri Unit 405, Office Block, Hotel Equatorial Shanghai
Slavka Krautzeka 83/A No.65, Yan An Road (West), Shanghai, 200040, China
51000 Rijeka, Croatia
Phone: +385 (51) 770 447 Phone: +86-21-62489820
Fax: +385 (51) 686 166 Fax: +86-21-62489821
www.intechopen.com

You might also like